Earth and Planetary Science Letters 286 (2009) 110–121 Contents lists available at ScienceDirect Earth and Planetary Science Letters j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / e p s l Late Paleocene high Laramide ranges in northeast Wyoming: Oxygen isotope study of ancient river water Majie Fan ⁎, David L. Dettman Department of Geosciences, University of Arizona, Tucson, Arizona 85721, United States a r t i c l e i n f o Article history: Received 26 September 2008 Received in revised form 16 June 2009 Accepted 16 June 2009 Available online 19 July 2009 Editor: T.M. Harrison Keywords: Rocky Mountains paleoelevation oxygen isotope ratios Laramide freshwater bivalve a b s t r a c t The distribution and initial timing of the establishment of high surface elevations in the Rocky Mountains during the Early Cenozoic remain controversial despite the importance of these data in testing tectonic models for this region. We track the timing and distribution of high elevation in the Rocky Mountains during the Late Cretaceous–Early Eocene by examining annual and seasonal δ18O values of the ancient river water, which are calculated from the δ18O values of well-preserved freshwater bivalve shells. In the Powder River basin of the eastern Laramide province, the δ18O values of the ancient river water vary between − 23.0‰ and − 8.0‰SMOW in both seasonal and annual records in the Late Paleocene–Early Eocene. The large variation suggests that the ancient rivers were fed yearly or seasonally by snowmelt from highlands of 4.5 ± 1.3 km. This can be explained by the existence of the Bighorn Mountains and Black Hills with a drainage pattern similar to the present in northeast Wyoming. The δ18O values of ancient river water along the front of the Sevier thrust belt generally follow a trend from lower values in north, −14.2 ± 1.4‰ in the Early Paleocene Crazy Mountains basin, to higher values in south, − 11.1 ± 0.8‰ in the Late Paleocene Bighorn basin, and − 7.1 ± 1.6‰ in the Early Eocene Washakie basin. The variations within each basin are relatively small. These rivers most likely rise in the Sevier thrust belt, and may reflect highland elevation of 1–2 km. The δ18O values in the Alberta foreland and Williston basin are very low (− 20.5‰) in the Late Cretaceous, indicating the rivers were fed by snowmelt from the Canadian Rocky Mountains of 4.3 ± 1.0 km high. The attainment of high elevation in the eastern Laramide province prior to the western province could be explained by southwestward progression of back-thrusts soled into an earlier east-directed master detachment, which may be formed by the westward rollback of subducted shallow slab. Published by Elsevier B.V. 1. Introduction The Laramide orogeny of the Rocky Mountains is a system of basement-cored uplifts and intervening basins that formed ~80– 40 Ma in the foreland basin of the Sevier thrust belt in the western United States (Dickinson and Snyder, 1978; Bird, 1998; DeCelles, 2004). Analogous to modern flat-slab subduction in western South America (Jordan and Allmendinger, 1986), Laramide uplifts are the result of the NE–SW compression due to shallow subduction of the Farallon Plate beneath the North American Plate (e.g., Dickinson and Snyder, 1978; Bird, 1998; Saleeby, 2003; DeCelles, 2004). Although this region is very well studied, fundamental questions remain unanswered: how does shallow subduction thicken foreland crust and produce a landscape with intervening basins and ranges, and how is basement-involved deformation connected to the thin-skinned Sevier fold and thrust belt. A range of tectonic models have been proposed to explain the mechanisms of Laramide deformation: basal traction (Bird, 1998), lithospheric buckling (Tikoff and Maxson, 2001), intracrustal flow ⁎ Corresponding author. Tel.: +1 520 621 6014; fax: +1 520 621 2672. E-mail address: [email protected] (M. Fan). 0012-821X/$ – see front matter. Published by Elsevier B.V. doi:10.1016/j.epsl.2009.06.024 (McQuarrie and Chase, 2001), and thrust, back-thrust and crustal detachment (Erslev, 1993, 2005). The timing, magnitude, and pattern of high surface topography in the Laramide province need to be examined to help answer these basic questions and to test tectonic models. Oxygen isotope ratios of precipitation derived from authigenic minerals (e.g., paleosol carbonate, biogenic apatite and aragonite) have been applied as paleoaltimeters in numerous studies (e.g., Dettman and Lohmann, 2000; Garzione et al., 2006; DeCelles et al., 2007; Quade et al., 2007). This approach is based on the decrease in oxygen isotope ratios of precipitation as elevation increases, which is controlled by progressive condensation of atmospheric water vapor due to cooling or adiabatic expansion as the vapor mass ascends, leading to Rayleigh isotope fractionation (Dansgaard, 1953; Rowley, 2007). At present, there have been only a few studies addressing the paleoelevation of the Laramide Rocky Mountains (Gregory and Chase, 1992; Norris et al., 1996; Wolfe et al., 1998; Dettman and Lohmann, 2000; Fricke, 2003; Sewall and Sloan, 2006). Even though these studies have been mostly focused on Eocene elevations, they have yielded highly varying results. By using leaf margin analyses, Wolfe et al. (1998), and Gregory and Chase (1992) suggested that the Eocene M. Fan, D.L. Dettman / Earth and Planetary Science Letters 286 (2009) 110–121 111 Fig. 1. Shaded relief map of the central and Canadian Rocky Mountains. Symbols denote sampling locations and mean δ18O values of fossil shells. Basins discussed in text are bordered by dotted line. AB: Alberta foreland basin; BH: Bighorn basin; BlH; Black Hills; BM: Bighorn Mountains; BT: Beartooth Mountains; CMB: Crazy Mountains basin; GR: Green River basin; PD: Powder River basin; SC: Sage Creek basin; WR: Wind River basin; WRR: Wind River Range; WaB: Washakie basin; WiB: Williston basin. Arrows are paleocurrent directions (Flores and Ethridge, 1985; Dickinson et al., 1988; DeCelles et al., 1991; Borrell and Hendrix, 2000; Mack and Cole, 2005). Fig. 2. Generalized stratigraphic columns of sedimentary successions involved in each studied basin (age determinations of strata with sampled fossil unionid are listed in Table 1). Key references: Cherven and Jacob, 1985; Flores and Ethridge, 1985; Lerbekmo and Coulter, 1985; Robinson and Honey, 1987; Dickinson et al., 1988; DeCelles et al., 1991; Eberth and Hamblin, 1993; Buckley, 1994. 112 M. Fan, D.L. Dettman / Earth and Planetary Science Letters 286 (2009) 110–121 Table 1 Age constraints for the study intervals in each basin. Sample location Formation Alberta foreland basin 1 Alberta foreland basin 2* Williston basin Oldman and Dinosaur Park Formations Horseshoe Canyon Formation Crazy Mountain basin Bighorn basin Bear, Lebo, and Melville Member of Fort Union Formation Fort Union and Willwood Formations Powder River basin Washakie basin Tongue River Member of Fort Union and Wasatch Formations Luman Tongue Member of Green River Formation Hell Creek Formation and Tullock Member of Fort Union Formation Age (Ma) Constraint 74–76 68–73 62.3–68 61–65 53.6–59 54–58 53–54 Dating method Bentonite bed above the Dinosaur Park–Oldman K–Ar, Ar–Ar Formation contact C31r.1r–C32r.1r Magnetostratigraphy K–T boundary, bentotite beds Interval of C26r–C29r, Reference Eberth and Hamblin (1993) Lerbekmo and Coulter (1985) (Swisher et al., 1993; Lund et al., Ar–Ar, Magnetostratigraphy, 2002) NALMA Magnetostratigraphy (Buckley, 1994; Butler et al., 1987) Interval of C24n–C26r, carbon isotope excursion Ar–Ar and Magnetostratigraphy Coal stratigraphy, occurrence of Platycarya NALMA, pollen , and correlation with Bighorn Basin Biostratigraphy Volcanic tuff above the Member, top of C24n Ar–Ar and Magnetostratigraphy (Koch et al., 1995; Secord et al., 2006) (Wolberg, 1979; Robinson and Honey, 1987; Wing et al., 1991) Smith et al. (2003) Note: *Only one shell bed in the locality. southern Rocky Mountains were more than 3 km high, similar to today's elevation. This agreed with the δ18O values of lake microbial carbonates in the Green River basin (Norris et al., 1996), the δ18O values of unaltered freshwater bivalves (Dettman and Lohmann, 2000), and regional GCM modeling of Early Paleogene Laramide foreland (Sewall and Sloan, 2006). However, Morrill and Koch (2002) concluded that diagenesis may have altered the δ18O of lacustrine microbial carbonates in the Green River basin. Moreover, Fricke (2003) argued that Early Eocene Laramide range elevations were ~500 m based on similar δ18O values of mammal teeth from three Wyoming basins. These conflicting conclusions may arise from a number of factors: 1). diagenetic overprinting of original isotopic patterns; 2). not accounting for all the non-altitude factors that can affect the δ18O of surface water (e.g., latitude, temperature); 3). small sample numbers failing to document a regional pattern in surface water δ18O values. In this study we survey the oxygen isotope composition of fossil freshwater bivalves (Unioniacea superfamily) of Late Cretaceous–Early Eocene age collected from six basins in the Laramide tectonic province. Unionids have relatively thick growth increments aiding the study of seasonal isotopic variation. Our study examines both the seasonal and annual δ18O values of ancient river waters as recorded in fossil shells. We compare these to the δ18O values of modern precipitation and river water in our discussion of the paleoelevation of ancient river catchments. We then track the timing and spatial patterns of high elevation regions in Laramide ranges. 2. Regional setting Fossil shells were collected from the Alberta foreland basin, western Williston basin, Crazy Mountains basin, northern Bighorn basin, Powder River basin and southern Washakie basin (Figs. 1 and 2). Prior to Laramide deformation, these regions were a broad foreland basin of the thin-skinned Sevier fold-thrust belt. Laramide deformation partitioned the central Rocky Mountain foreland into discrete local basins separated by basement-cored uplifts with NE–SW to E–W orientations (Dickinson et al., 1988). The uplifts are bounded by moderately dipping to high-angle faults or are broadly anticlinal. The Crazy Mountains basin, Bighorn basin, Powder River basin and Washakie basin are of this type. Late Cretaceous through Miocene age sediments derived from the Sevier thrust belt and Laramide uplifts were deposited in these basins. Fossil shells from the Washakie basin are from the Luman Tongue Member of the Green River Formation, a Fig. 3. Measured and modeled δ18O values of ancient river water for western interior of North America plotted against paleo-latitude of sampling location. Dotted lines are the 1σ uncertainties of modeled low elevation δ18O values based on paleolatitude data in Table 2 and corrections discussed in text. See text for details of calculations and modeling. M. Fan, D.L. Dettman / Earth and Planetary Science Letters 286 (2009) 110–121 113 Fig. 4. Seasonal δ18O variation of representative fossil shells in the studied basins. A: Powder River basin; B: western Williston basin; C: Alberta foreland basin; D: Crazy Mountains basin; E: Bighorn basin; F: southern Washakie basin. mix of fluvial and lake facies in the early stages of Lake Gosiute (Sklenar and Anderson, 1985). Shell samples in the other three basins are from Paleocene to Early Eocene fluvial sediments. In the Alberta foreland basin, Laramide age deformation is an eastward continuation of thinskinned Sevier fold and thrust, which overthrusted Mesozoic shale and molasse, but did not generate any basement uplifts (Obsborn et al., 2006). Shells from the Alberta foreland basin were collected from sediments of Late Campanian and Maastrichtian age. In the Williston basin, a depression in the Canadian Shield, fossil shells were collected from Late Cretaceous and Early Paleocene fluvial sediments. Sample ages are based on radiometric ages, paleomagnetic ages and intrabasin mammalian biostratigraphic correlation with magnetostratigraphy (Table 1). 3. Analytical methods Aragonite shells used in this study are all unaltered as judged by physical appearance, cathodoluminescence microscopy and X-ray diffraction for a subset of the samples. X-ray diffraction was performed with a Bruker D8 Advance Diffractometer using Cu Kα radiation. When possible, ten shells and shell fragments from each stratigraphic horizon were analyzed for stable isotopes. The bulk shell δ18O values were mostly presented in Dettman and Lohmann (2000), with new analyses added (supplementary data). Total of 881 individual shell fossil analyses are included in this paper. Shell Aragonites were collected by drilling through the shell body, integrating isotopic variation and yeilding a growth amount-weighted average δ18O value. The bulk δ18O value presented in this paper refers to the average values of all analyzed bulk shell samples in stratigraphic horizon except the Powder River basin, where some data represent single shell analyses. In addition, we micro-milled 12 selected samples in order to study the seasonal isotopic variation of the Early Cenozoic rivers. Shells were sectioned along the axis of maximum growth, mounted as thick sections (~ 1 mm). Growth bands in cross-sectioned shells were subsampled using a computer-controlled micro-mill with a 30 μm sampling resolution (Dettman and Lohmann, 1995). The δ18O and δ13C values of aragonites were measured using an automated carbonate preparation device (KIEL-III) coupled to a gas-ratio mass spectrometer 114 M. Fan, D.L. Dettman / Earth and Planetary Science Letters 286 (2009) 110–121 Seasonal isotopic variation of selected shells from each basin are used to compare ancient and modern river seasonal variability in the Laramide region (Figs. 4 and 5). Oxygen isotope ratios of modern river water are from Coplen and Kendall (2000), modern precipitation are from the United States Network for Isotopes in Precipitation (USNIP) (www.uaa.alaska.edu/enri/usnip), Canadian network for Isotopes in Precipitation (CNIP) (www.science.uwaterloo.ca/~twdedwar/cnip), and Dutton et al. (2005). 5. Calculation of ancient river water δ18O values The oxygen isotope composition of Unionid shell aragonite is controlled by the temperature and δ18O value of the river water in which it grew (Grossman and Ku, 1986; Dettman et al., 1999). Here, we use the empirically determined relationship between the bulk δ18O values of Unionid shell and mean annual river water in temperate climates from Kohn and Dettman (2007): 18 18 2 δ Oðshell;PDBÞ = ð0:892 0:024Þ δ Oðriver water;SMOWÞ 0:978 0:240ðR = 0:98Þ ð1Þ Fig. 5. Seasonal δ18O variation (black circles) of three kinds of river to the east of Rocky Mountains. A: Milk River; B: Halfmoon Creek; C: Big Sioux River. River water δ18O data are from Coplen and Kendall (2000). Temperature (gray circles) is the average temperature for the day of sampling from a nearby climate station (NCDC: http://www. ncdc.noaa.gov/oa/ncdc.html). to calculate the mean annual δ18O values of ancient river water. Because these freshwater bivalves stop growing below approximately 10–12 °C and their growth is heavily biased to late spring and early summer temperatures, a large majority of shell aragonite is produced in a limited range of temperatures (20–25 °C) (Dettman et al., 1999). This growth temperature bias leads to a good linear correlation between mean annual river water and bulk shell δ18O values, which can be used to calculate the mean δ18O value of ancient river water if the climate is temperate and seasonal (Kohn and Dettman, 2007). If temperatures were not seasonal, but remained at one temperature extreme (e.g.12° or 30°) throughout the year, then calculated river water δ18O values could be as much as 2‰ too high or too low, but this seems extremely unlikely given prominent growth bands in the shells, indicating seasonal growth cessation, and the botanical and modeling evidence for moderate seasonality (Wilf, 2000; Sewall and Sloan, 2006). Our calculated mean annual δ18O values of ancient river waters range from −23‰ to −5‰SMOW (Fig. 3). Each shell sample represents an average of several years of growth in these river systems. Although the average river water δ18O values calculated from a single shell could be affected by a few years of (Finnigan MAT 252). Samples (20 to 150 μg) were reacted with dehydrated phosphoric acid under vacuum at 70 °C. The isotope ratios are calibrated based on measurements of NBS-19 and NBS-18; precision is ±0.1‰ for δ18O and ±0.06‰ for δ13C (1σ). 4. Oxygen isotope results The δ18O values of bulk shells from basins along the front of the Sevier thrust belt follow a trend from lower values in north (−13.7 ± 1.3‰PDB, Crazy Mountains basin) to higher values in south (−10.7 ± 0.8‰PDB, Bighorn basin, and −7.3 ± 1.5‰PDB, Washakie basin), and the within-basin variation is small. In the Powder River basin and western Williston basin, shell δ18O values show large variation between −22.9‰ and −7.9‰ PDB, and −21.8‰ and −9.3‰PDB, respectively. Our results in the western Williston basin are consistent with values from the eastern Williston basin (Carpenter et al., 2003; Cochran et al., 2003). The δ18O values of shells in the Alberta foreland basin vary between −19.3‰ and −13.3‰ PDB (Fig. 3). Fig. 6. Regression for sampling station latitude and the δ18O values of river water in lowelevation stations (b200 m) within the USA (data from Coplen and Kendall, 2000). Open circles are rivers in Hawaii (b23° latitude) or rivers that are sourced in high elevation catchments with sampling stations at low-elevation, which are not included in the regression. M. Fan, D.L. Dettman / Earth and Planetary Science Letters 286 (2009) 110–121 115 Table 2 Paleomagnetic data and paleolatitude. Sample location Age (Ma) Present latitude (Nº) Present longitude (Wº) Pole (Nº) Pole (Eº) A95 P Paleolatitude (Nº) Southward transport (º) Alberta foreland basin 1 Alberta foreland basin 2 Williston basin Crazy Mountains basin Bighorn basin Powder River basin Washakie basin 74–76 68–73 68–62 65–61 54–59 54–58 53–54 49.16 50.95 47.55 46.21 44.84 44.8 40.89 110.41 112.09 107.01 109.73 108.81 106.08 108.66 74.70 74.20 74.20 75.90 75.90 75.90 77.90 204.70 204.80 204.80 196.80 196.80 196.80 179.30 5.90 3.20 3.20 2.90 2.90 2.90 3.40 0.55 0.51 0.59 0.64 0.67 0.68 0.81 58.43 60.75 56.36 53.18 51.67 51.01 43.55 9.3 ± 5.9 9.8 ± 3.2 8.8 ± 3.2 7.0 ± 2.9 6.8 ± 2.9 6.2 ± 2.9 2.7 ± 3.4 Note: A95 is 95% confidence limits. P is the angular distance between sample site and reference pole (Eq.A20 in Butler, 1992). Paleolatitude is calculated from Eq. A26 in Butler (1992). Southward transport is the latitude difference (and 95% confidence limits) between present latitude and paleolatitude. anomalous precipitation patterns, the chances of this are reduced by averaging ten shells from each stratigraphic level. 6. Constraints, corrections, and application to paleoaltimetry Knowing the δ18O values of ancient river water provides some insight into the paleoelevation of ancient river catchments, but many other factors can affect the δ18O of rainfall. Many of these effects (e.g. temperature, continentality) are combined into the strong relationship between the δ18O of rainfall and latitude (Dutton et al., 2005). Therefore we will attempt to remove the latitudinal effect on δ18O values by comparing the ancient river δ18O values to modern low-elevation rivers (representing low-elevation rainfall δ18O values) across a range of latitudes. We use river water δ18O data because it is much more abundant than precipitation data and it comes closer to a weighted average of the δ18O of precipitation in the river's catchment. Accounting for latitudinal differences is particularly important because North America moved southward, changing latitude by about 10°, during the time interval under study, and the study region spans another 10° of latitude. This low-elevation data set will be modified based on studies of Paleogene meteoric water δ18O values to make the comparison more appropriate for Late Cretaceous and Cenozoic times. Our data also provides a check on the latitudinal gradient of δ18O values for low elevation rainfall in the late Cretaceous and Paleogene. The δ18O value of modern low elevation (b200 m) river water across the USA is related to the latitude of river stations (Fig. 6): 18 δ Oðriver water;SMOWÞ = ð0:0012 0:0009Þ j LATj 2 ð0:3940 0:0816Þ jLATj ð2Þ 2 + 10:2154 1:7811 ðR = 0:91Þ In the dataset of Coplen and Kendall (2000), the elevations of the modern river water samples in low elevation (b200 m) are not necessarily the elevations of river source catchments. Many rivers have significantly higher source catchments elevation than sampling site, like Columbia River sampled in Washington, and Missouri River sampled in South Dakota and North Dakota. Therefore, in our regression we do not include the rivers that clearly rise at high elevation, and Hawaiian rivers, which are not applicable to continental interiors. This significantly improves the regression in Dutton et al. (2005). The key factors affecting the δ18O values of either modern or ancient river waters include water vapor source, vapor temperature, source water δ18O values, latitude, and elevation. The effect of evaporation on the δ18O value of river water is usually quite small, particularly in humid climates, and is ignored in this paper. The modern climate of the Rocky Mountain region and the western Great Plains is influenced by the competition of three air masses that originate over the Arctic, the Gulf of Mexico, and the Pacific Ocean (Bryson and Hare, 1974). The Rocky Mountains, together with the Basin and Range to the west, form a large high topographic barrier leading to a rain shadow that reduces the contribution of Pacific-sourced moisture to the Laramide Rocky Mountain province. To the east of the Rocky Mountains, moisture from the Gulf of Mexico brings relatively abundant summer rainfall. Prior to the Latest Cretaceous–Paleocene, the movement of Sevier fold-andthrust belt caused significant shortening and thickening, forming a highelevation hinterland plateau with rugged topographic front to the west of the Rocky Mountains (DeCelles, 2004; DeCelles and Coogan, 2006), Given the topographic similarity to today, we assume that large-scale climate pattern during Early Cenozoic was similar to present and most of the precipitation is sourced from the Gulf of Mexico region. Temperature of the sea surface and middle latitude continental interior in the Late Cretaceous–Early Eocene was higher than present (Zachos et al., 1994; Wilf, 2000; Zachos et al., 2001). Higher temperature can influence the δ18O values of both source region water vapor and water condensates (Dansgaard, 1953), which could lead to a different stable isotope–latitude relationship than that of today. One study of the Early Eocene latitudinal-isotopic gradient in river water shows that the gradient was similar to today in the middle latitudes, although there is a significant difference in the intercept of this relationship (Fricke, 2003). The North American Plate has moved southward at least 5° since the Late Cretaceous (Besse and Courtillot, 2002), which could result in a ~2.4‰ difference between the δ18O values of precipitation at present and in the Late Cretaceous at a location (Fig. 6, and Dutton et al., 2005). Neglecting this paleolatitudinal correction will lead to overestimates of the paleoaltitude by at least 0.8 km (if a lapse rate of 2.8‰/km was applied to this difference; Poage and Chamberlain, 2001). The paleolatitude of our sample localities are derived by using the paleopole locations in Besse and Courtillot (2002), and the dipole equation in Butler (1992). The lower and upper limits of paleolatitude are derived from the confidence limits of the paleopole (Α95) (Table 2). In order to discuss ancient river water δ18O values in terms of stable isotope patterns seen on the globe today, we use the regression relationship between the δ18O value of modern low elevation river water and sampling latitude (Fig. 6), incorporate corrections for 1) higher intercept due to warmer temperature in the Late Cretaceous– Early Eocene (+4 ± 2.8‰, Fricke, 2003); 2) changes in the δ18O value of seawater due to smaller global ice volume (−1‰, Zachos et al., 1994); and 3) latitude change due to continental drift of North America since the Late Cretaceous–Early Eocene (Besse and Courtillot, 2002), to model the low elevation river water 18O values. Errors propagated from the three corrections yield a δ18O range for low elevation river water (Fig. 3). If an ancient river catchment was at low elevation, the determined δ18O values of ancient river water should be in the range of these modeled δ18O values. Note in Fig. 3 that our most positive δ18O values, probably sampling the lowest elevation catchments, are within the ancient lowelevation river water field, suggesting that the corrections approximate low-elevation precipitation relatively well. If the ancient river rises from a high elevation catchment, the determined δ18O values should be lower than the area of modeled δ18O values. The elevation of the river source terrane can then be calculated from the difference in the δ18O value between the ancient river water and the lower limit of the modeled low elevation river water, and a lapse rate of −2.8(±0.6)‰/km. This leads to a conservative estimate of elevation as we use the lower limit of the 116 M. Fan, D.L. Dettman / Earth and Planetary Science Letters 286 (2009) 110–121 modeled low elevation river water. This lapse rate is derived from global precipitation δ18O values (Poage and Chamberlain, 2001), and is similar to the lapse rate derived from North America precipitation (−2.9‰/km Dutton et al., 2005). Although our bivalve-based data reconstructs river water δ18O values, we use an isotopic lapse rate based on precipitation instead of one based on river water. This is because we will discuss the elevation of the catchments of these ancient river systems rather than the elevation of the point in the river system sampled by the bivalves. In general river sample isotopic lapse rates are much higher than precipitation-based ones because the former reflect rainfall in the river catchment at higher elevations than that of the sampling point (e.g. Dutton et al., 2005). If a river has a high elevation catchment the δ18O values of the river water will tend to be lower, even if the sample of river water is collected near sea level. The biggest problem with using the river water lapse rate is the lack of information on the elevation difference in a river's catchment and that the elevation of a single sampling point in a river system has little to do with the overall elevation of its catchment (Kohn and Dettman, 2007). The shells in this study are limited to the basin floors, where sediment accumulates and preserves these fossils. In some basins there is a very large range in the δ18O of river waters, a range that clearly represents variation in the elevation of the different river catchments, and not dramatic changes in the elevation of the basin floor (Dettman and Lohmann, 2000). The Western Interior Seaway covered a large area of North America in the Late Cretaceous (Roberts and Kirschbaum, 1995), and this water body probably dominated the water vapor cycle of the Campanian and Maastrichtian Alberta foreland and Williston basins. Because this was basically a coastal system we will not use the low elevation latitude– rainfall δ18O relationship of Fig. 3, which is more appropriate for a continental interior region, to interpret the Alberta and Williston basin data. In this case, we will use a modern empirically-tested model of the Δδ18O–elevation relationship to reconstruct river catchment elevations (Rowley et al., 2001; Rowley, 2007). The Δδ18O value is the δ18O difference between highland precipitation and lowland precipitation. The modeled Δδ18O-elevation relationship, which tracks a single rising package of water vapor cannot be applied to the Early Cenozoic samples because of the retreat of the interior seaway and the possibility that water vapor came into the Laramide region from different sources. The uncertainties reported in this paper, encompassing the model, corrections, and calculation of paleoelvation all are one standard deviation (1σ). 7. Seasonal δ18O variation in ancient and modern river waters The δ18O values of river waters are controlled by surface runoff and groundwater in river catchments. Rivers with low elevation catchments tend to have higher δ18O values in contrast to high elevation catchments. High-elevation snowmelt can feed lowland rivers as surface runoff during the spring and early summer or year-round as groundwater. The δ18O values of this type of river water is often very low in spring and early summer, equivalent to the δ18O values of high-elevation snowmelt (Horton et al., 1999). At present, in the Laramide region, summer precipitation amount is about two times that of winter precipitation amount (Dutton et al., 2005), and the highest river discharge is generally in March–June (SAGE River Discharge Database, www.sage.wisc.edu/ riverdata). We group the isotopic patterns of rivers in this region into three types. 1. Rivers originating at high elevation with different sources of water seasonally. The Milk River flows east from Glacier National Park, Montana, with catchment elevations between 1.0 and 2.8 km. Precipitation at the closest rain monitoring station (MT-05, Glacial National Park, 968 m) averages −18.3‰SMOW with a range of 15.6‰ (Dutton et al., 2005; USNIP). The large seasonal δ18O variation in the Milk River at the Nashua station (48.13°N, 106.36°W, 696 m) shows that the river is fed by snow melt from elevations higher than station MT−05 in spring with δ18O values as low as −23‰. In the summer, local and lower elevation rain seems to dominate, with δ18O value as high as −10‰. During the winter, when surface runoff is frozen, groundwater supports the river with intermediate δ18O values (Fig. 5A). 2. Rivers with year-round high-elevation sources of precipitation. Halfmoon Creek rises in the Sawatch Range, central Colorado, which contains several of the highest peaks (N4 km) in the Rockies. At station Malta (39.17°N 106.39°W, 2996 m), the δ18O value is low and very stable throughout the year, −17.0 ± 1.2‰. There is a small increase in the δ18O value of river water in the late summer (Fig. 5B). Precipitation at the closest station (CO-02, Niwot Saddle, 3520 m) has an annual average δ18O value of −14.3‰ (USNIP). This suggests that the river is primarily fed by groundwater that integrates annual precipitation, perhaps derived from higher elevations. Summer local precipitation seems to increase river water δ18O values only slightly. 3. Rivers originating at low elevation. The Big Sioux River catchment begins in Roberts County, South Dakota, with an elevation less than 500 m. At station Akron (42.83°N 96.56°W, 341 m), the seasonal δ18O variation of the river water is very small, −8.6±1.0‰ (Fig. 5C). The annual average δ18O value of precipitation at the closest station (MN27, Lamberton, 343 m) is −8.6‰ (USNIP). The small seasonal isotopic cycle and similarity in δ18O value to precipitation suggests that the river is fed by ground water, which is recharged by local precipitation. Micromilled shells from the different areas surveyed show a variety of seasonal oxygen isotopic patterns: three types in the Powder River and Williston basins, two patterns in the Alberta foreland basin and only one pattern in the Crazy Mountains, Bighorn and Washakie basins (Fig. 4). Although the amplitude of seasonal oxygen isotope cycles in shells are not directly comparable to river water δ18O cycles due to the seasonal temperature variation, large changes in river water δ18O values (greater than 4‰ in many cases) must be reflected in the δ18O cycles in shells. This is because the limited range of shell growth temperatures can only cause ~4‰ variation in shell δ18O values (Dettman et al., 1999). In the Powder River basin, shell PR10-86A has a large amplitude δ18O cycle (−17.9‰ to −7.0‰), similar to that of the Milk River. This highamplitude shell must have lived in a river with large seasonal δ18O amplitude, suggesting that the river was seasonally fed by snowmelt from high elevation either as groundwater or as surface runoff. Powder River basin shell PR2-85 has an isotopic cycle similar to the Big Sioux River. It has small δ18O amplitude (between −11.8‰ and −7.0‰) and the mean δ18O value is relatively high. This kind of river is found today at low elevations, fed by local precipitation. PR3-83 is similar to Halfmoon Creek, with a small δ18O amplitude, but with a very low mean δ18O value (−24.3 ± 2.6‰). This kind of river usually originates in high elevations and is fed by a groundwater supply that is recharged at high elevation (Fig. 4A). In the western Williston basin, shells with these three types of seasonal δ18O patterns are present (Fig. 4B). In the Alberta foreland basin, seasonal δ18O variation in ancient river water recorded in samples 93-13 and Tyrell, is very similar to Big Sioux River, with a small amplitude and higher δ18O values (−13.6± 1.0‰). In contrast, one shell, sample 93-11, is similar to Halfmoon Creek, also low amplitude, but with very low δ18O values (−18.9 ± 0.8‰) (Fig. 4C). In the Crazy Mountains basin, Bighorn basin, and Washakie basin, the seasonal δ18O variation in shells have relatively small amplitudes and are intermediate in average δ18O value. These rivers are of Big Sioux River type, and fed by local precipitation (Fig. 4D, E, and F). Our Washakie basin shell data are similar to seasonal records from other well preserved shells from the Green River basin (Morrill and Koch, 2002). 8. Discussion 8.1. Late Cretaceous High Canadian Rocky Mountains The δ18O value of paleosol carbonate in the Late Maastrichtian lower Willow Creek Formation in southern Alberta is −12.1 ± 0.9‰PDB M. Fan, D.L. Dettman / Earth and Planetary Science Letters 286 (2009) 110–121 117 was active during the Late Cretaceous, and that a high orogenic plateau sat to the west of the basin (Van der Pluijm et al., 2006). The δ18O range of river water in the Williston basin is − 18.3 to −21.5‰SMOW. This is 8.5 to 11.7‰ lower than the δ18O value of the low elevation river water (Fig. 3) and is similar to that of the Alberta foreland basin. This suggests mountains with similar elevations, in this case 4.3 ± 1.0 km, were present (Fig. 7A). Oxygen isotope cycles in ancient river water further suggest that highland snowmelt fed some rivers seasonally or yearly (Fig. 4B). Rivers rising in the lowlands of the basin have much higher δ18O values that fall within the range of modeled low elevation river water (Fig. 3) and seasonal records of shell δ18O values reflect both high and low elevation catchments. The upper Cretaceous and lowest Cenozoic strata in southeastern Montana contains high amounts of volcanic lithics and andesitic conglomerate, and paleoflow directions are from the southwest to west in the region; both suggest that the sediment source terrane was the Sevier thrust belt in Idaho and southwest Montana (Cherven and Jacob, 1985; Fastovsky, 1987). Therefore, the high elevation catchments for rivers in the western Williston basin were most likely in the Rocky Mountains. Large rivers frequently carry high elevation snowmelt N500 km without any significant change in the water δ18O value and this is the most plausible explanation for the low δ18O values in Williston basin (Fig. 8A). Examples of rivers carrying water with very negative δ18O values over long distances are the Colorado and Missouri Rivers, which have significantly lower δ18O values in Arizona and South Dakota than the local precipitation (Coplen and Kendall, 2000). 8.2. Late Paleocene High surface elevation of eastern Laramide ranges Fig. 7. A): Estimated paleoelevation of the Late Cretaceous and Early Paloecene Canadian Rocky Mountains by using thermodynamic based lapse rate in Rowley, 2007; B): Estimated paleoelevation of mid Paleocene-Early Eocene river source region by using lapse rate against present latitude of sampling sites in the Laramide Rocky Mountain basins. Elevation of river catchment with measured δ18O values higher than modeled low-elevation values are arbitrarily assigned as zero. (n = 7) (Mack and Cole, 2005). From this we calculate the δ18O value of lowland mean annual precipitation as −10.3 ± 2.5‰ using the calcite oxygen isotope fractionation relationship (Kim and O'Neil, 1997) and a temperature of 12–30 °C, assumed to be the formation temperature of soil carbonate during warm growing season (Cerling and Quade,1993). Although evaporation can increase the δ18O values of soil water relative to mean annual precipitation, this effect is minimal in wet climates (Quade et al., 2007). Given the proximity of the interior seaway and wet climate in the Late Cretaceous (Roberts and Kirschbaum, 1995; Golovneva, 2000), strong evaporation is very unlikely. The similar δ18O values between the lowland precipitation and the highest river water suggest that the shells with the relatively positive oxygen isotope cycle in Fig. 4C lived in river water fed mainly by local precipitation. The occurrence of much lower values can therefore be attributed to the low δ18O values of distal highland precipitation. This can be seen in the shell 93-11 in Fig. 4C, which grew in water that had low δ18O values all year long. Numerous other shells with bulk δ18O values in the −18‰ to − 20‰ range have similar cycles and these low values cannot be the result of seasonal low-elevation snowmelt events. Our estimated highland elevation is 4.3 ± 1.0 km during Late Cretaceous (Fig. 7A). In the Alberta foreland basin, paleocurrent directions are generally east to southeast (Mack and Cole, 2005), suggesting that high elevation snowmelt came from the Canadian Rocky Mountains to the west of the basin. This is consistent with tectonic reconstructions of this region, which argue that thrusting In the Powder River basin, the lowest δ18O value of the river water is −22‰SMOW, which is ~ 11‰ below the minimum δ18O value modeled for low-elevation ancient river water (Fig. 3). The negative offset reflects high elevation in the river catchments. Seasonal oxygen isotope variation in ancient shells further suggests the presence of both highland and low elevation precipitation in local rivers (Fig. 4A). The estimated elevation of the mountain ranges that provide runoff to the Powder River basin is 4.5 ± 1.3 km in the Late Paleocene (Fig. 7B). The simplest scenario for the paleography in the Powder River basin based on these three different types of river isotopic records would be a variety of river catchments that carried distal high elevation waters and/or basinal lowland precipitation (Fig. 8B). The high elevation region in the eastern Laramide province was most likely the Bighorn Mountains in northern Wyoming and Black Hills in South Dakota. Today the rivers of this region flow east and northeast from the high Laramide ranges in northern Wyoming and southwest Montana. These rivers have low δ18O values due to high elevation snowmelt input. For example, the Tongue River which flows northward from the northern Bighorn Mountains has a δ18O value of −16.6 ± 2.0‰SMOW; tributaries of the Powder River with δ18O values of ~−17‰ flow from the eastern Bighorn Mountains and western Black Hills toward the center of the basin merging as the Powder River (Coplen and Kendall, 2000). The Powder River and Tongue River then join the Yellowstone River in southwestern Montana, and form a large tributary of Missouri River. A Late Paleocene scenario similar to the modern drainage pattern could explain the very low δ18O values of the river waters in the Powder River basin. Such a paleodrainage system is supported by the paleogeographical reconstruction of the Powder River basin, which is very similar to the pattern of the modern drainage system just described (Flores and Ethridge, 1985). Low-temperature thermochronology studies also support this paleogeography, suggesting major exhumation of the Bighorn Mountains at 65 ± 5 Ma and Early Paleocene cooling of the Black Hills (Strecker,1996; Crowley et al., 2002). In addition, Late Paleocene–Early Eocene synorogenic conglomerates with a high proportion of Precambrian basement clasts are found along the eastern flank of the Bighorn Mountains (Hoy and Ridgeway, 1997). 118 M. Fan, D.L. Dettman / Earth and Planetary Science Letters 286 (2009) 110–121 Fig. 8. Paleodrainage reconstruction of the studied area based on the presented data, and the inferred elevation of the Rocky Mountains. Lines with arrows stand for the paleorivers with the δ18O value shown. See Fig. 1 for names of major structures and basins, and references of paleoflow directions. Fricke (2003) pointed out that seasonal or episodic Arctic air mass incursions could result in very low δ18O values in the Powder River basin. Although this phenomenon can explain very negative δ18O values in seasonal precipitation (snow) at low elevations today, it is very unlikely to affect the average annual or warm season δ18O value of river waters. In southern Saskatchewan, where the present latitude is equivalent to the paleolatitude of the Williston basin and Powder River basin, winter snow δ18O values are often less than −25‰ (CNIP), but the modeled δ18O value of low-elevation river water is −16 ± 2‰, which is the same as the δ18O value of the local mean annual precipitation (CNIP). In this case the low-elevation rivers are not biased in favor of winter precipitation, rather they are supported by local groundwater. This argument is also supported by the fact that low elevation rivers in Minnesota and Michigan have δ18O values similar to that of the local mean annual precipitation, which is relatively high (−8 to −10‰), even though these areas are equally subject to significant water input from arctic air masses (Coplen and Kendall, 2000; USNIP). 8.3. Lower surface elevation of the western Laramide province before the Early Eocene The δ18O values of Paleocene and Early Eocene river water in the Crazy Mountains basin, Bighorn basin, and Washakie basin fall within or close to the range of modeled low elevation river water (Fig. 3), which suggests that the catchments of these ancient rivers were at relatively low elevation. Seasonal oxygen isotopic variation in shells from the three basins further shows that river water did not have extremely low δ18O values (b−15‰) seasonally or annually (Fig. 4D, E, and F). Although we suggest that this indicates low elevation catchments for the sampled rivers in these localities, we can not completely rule out some effect of local relief on these samples for two reasons: 1) both modeled and measured low elevation river water δ18O values have a range of ~4.5‰, which could mask ~2 km of elevation variability; 2) the δ18O values of ancient mean annual precipitation recorded in the paleosol carbonate nodules of Paleocene–Early Eocene age in this area are higher than our calculated river water δ18O values, which suggests an elevation difference between local and distal precipitation sources. The δ18O values of paleosol carbonates in the Wind River basin are −10.0‰ to −8.0‰PDB in the Early Eocene (our unpublished data), and are −10.0‰ to −7.5‰PDB in the Bighorn basin in the Late Paleocene–Early Eocene (Koch et al., 1995). Using a broad temperature range (12–30 °C) we can estimate the δ18O values of rainfall in these basins as between −4.5 and −11‰SMOW. Unionid shells in Bighorn basin (−12.6‰ to −8.6‰PDB) imply that water δ18O values are in the −8.5 to −13‰SMOW range (using Eq. (1), above). In the Sage Creek basin, southwest Montana, the δ18O values of paleosol carbonates varies between − 9.0‰ and −5.0‰PDB during Early Eocene Climatic Optimum, while the δ18O values of Paleocene fluvial calcite cements are −11.0‰ to −9.0‰PDB (Kent-Corson et al., 2006), again this suggests that river water δ18O values are lower than local precipitation. This difference is increased if the Eocene soils were warmer than the Paleocene rivers. The differences between the δ18O values of ancient river carbonates and soil carbonate in these basins suggest a maximum local relief of 1–2 km. Such relief is tectonically very plausible because crustal shortening and thickening of the Sevier fold-thrust belt during the Late Cretaceous–Paleocene formed a high-elevation hinterland plateau to the west of the Rocky Mountains (DeCelles, 2004; DeCelles and Coogan, 2006). Paleocene paleocurrent directions in the studied strata of the Bighorn and Crazy Mountains basins are generally eastward, which supports the scenario that rivers sourced in the Sevier thrust belt delivered water with lower δ18O values to the basins (Dickinson et al., 1988; DeCelles et al., 1991; Borrell and Hendrix, 2000). There are no robust paleocurrent direction measurements in the Luman Tongue Member in the Washakie basin. Our data therefore suggest there was no Laramide range in the western Laramide province as high as those in the eastern region in the Late Paleocene. M. Fan, D.L. Dettman / Earth and Planetary Science Letters 286 (2009) 110–121 119 Fig. 9. Schematic cross-sections showing the mechanism of forming high ranges in the eastern Laramide province earlier than the western province. Not to scale. 8.4. Implications This study adds to a growing body of knowledge regarding the topographic development of the Rocky Mountains. High elevation, 4.5 ± 1.3 km, developed in the eastern Laramide province, and a catchment draining the high regions of the Bighorn Mountains and the Black Hills formed in the Powder River basin by the Late Paleocene. Our data from the western Laramide province (Crazy Mountains, Bighorn, and Washakie basins) suggest that a maximum of 1–2 km of local relief existed between the Sevier thrust belt or Laramide ranges and basins in the Paleocene–Early Eocene. Therefore, ranges in eastern Laramide province reached high surface elevation earlier than ones in west. This pattern is analogous to the westward migration of the exhumation front in the Sierras Pampeanas and the high surface elevation of the Sierra Grande in the eastern Sierras Pampeanas (Coughlin et al., 1998). Significant elevation increase of Laramide ranges after the Late Paleocene in the western Laramide province is required to form the present landscape of the Rocky Mountains. Sedimentological and structural evidence have shown that Laramide deformation was initiated in the Late Cretaceous (e.g., Gries, 1983 and ref. therein, Dickinson et al., 1988; Bird, 1998; DeCelles, 2004). However, this study shows that Laramide ranges with elevations comparable to today's were not formed until the Late Paleocene in the eastern Laramide province, and that ranges in eastern province attained high topography prior to the western province. Erslev (1993) suggested that the Laramide uplifts are a system of thrusts and back-thrusts which were connected to a detachment in the lower curst. We therefore propose that the pattern we observe could be explained by southwestward progression of backthrusts soled into an east-directed master detachment that formed earlier (Fig. 9) (Erslev, 1993, 2005). The shallow subduction of the Farallon Plate beneath North America buckled the lithosphere in the Sevier foreland by basal traction or horizontal endloading of highlands in the Sevier hinterland during the Late Cretaceous to mid Paleocene time (Bird, 1998; Tikoff and Maxson, 2001). This buckling could have led to high Laramide ranges relative to the basins. The subsequent development of an east-directed crustal detachment may have formed the Black Hills and Bighorn Mountain uplifts in the Late Paleocene at the detachment tips. Following this, back-thrusts could develop in the folded area and lead to high ranges in the western Laramide province (Erslev, 1993, 2005). The development of the detachment thrust and back-thrusts may follow the magmatic sweep caused by eastward shallow subduction and westward rollback of the subducted slab (Constenius, 1996). Although our research is based on what is by far the largest paleosurface water δ18O data set collected to date, with good age control and multiple basin distribution, it is clear that a much more detailed temporal and spatial data set is needed to reconstruct the elevation history of the Laramide ranges and test our proposed kinematic model. In addition, more work is needed on the post-Laramide basins to document later erosional exhumation, uplift, or subsidence that led to the topographic patterns we see today. 9. Conclusion After removing the latitudinal δ18O effects on the δ18O of ancient river water, we can discuss the patterns of elevation change across the Late Cretaceous–Early Eocene foreland basin of the Canadian Rocky Mountains and the intermontane basins of the central Laramide Rocky Mountains. The Canadian Rocky Mountains were 4.3±1.0 km high during the Late Cretaceous, and snowmelt from the Rocky Mountains recharged the river water in the Alberta foreland basin, and western Williston basin both annually and seasonally. Laramide uplifts attained elevations up to 4.5± 1.3 km in northeast Wyoming in the Late Paleocene, and the most likely locus of this high elevation was the ancestral Bighorn Mountains and the 120 M. Fan, D.L. Dettman / Earth and Planetary Science Letters 286 (2009) 110–121 Black Hills. Paleodrainage patterns similar to the present developed in the Powder River basin in the Late Paleocene; rivers in the basin had catchments at a variety of elevations, some carrying high elevation precipitation, others sourced at low elevations. No comparably high topography developed in the western Laramide province before the Early Eocene. This east–west difference in timing of high elevation development for the Laramide ranges can be explained by southwestward progression of back-thrusts soled into an earlier northeast-directed master detachment, which may be formed by the westward rollback of a subducted shallow slab. Acknowledgements The bivalve fossil collections in this paper are largely based on the field notes of the late John H. Hanley. We would like to thank Jay Quade, Peter DeCelles and Paul Kapp for their helpful discussion. This manuscript was significantly improved by reviews by Carmala Garzione, and an anonymous reviewer. Appendix A. Supplementary data Supplementary data associated with this article can be found, in the online version, at doi:10.1016/j.epsl.2009.06.024. References Besse, J., Courtillot, V., 2002. Apparent and true polar wander and the geometry of the geomagnetic field over the last 200 Myr. J. Geophys. Res. 107, 2300–2331. Bird, P., 1998. Kinematic history of the Laramide orogeny in latitudes 35°–49°N, western United States. Tectonics 17, 780–801. Borrell, J.K., Hendrix, M.S., 2000. Cretaceous-Tertiary tectonic evolution of the northern Crazy Mountains Basin, Central Montana: a petrofacies, lithofacies, and paleocurrent approach. GSA Rocky Mountain Section Meeting. Bryson, R.A., Hare, F.K., 1974. The climates of North America. In: Bryson, R.A., Hare, F.K. (Eds.), Climates of North America. In: World survey of climatology, vol. 11. Elsevier, New York, pp. 1–14. Buckley, G. A., 1994. Paleontology, geology and chronostratigraphy of Simpson Quarry (Early Paleocene), Bear Formation, Crazy Mountains Basin, south-central Montana. Ph. D. Dissertation, Department of Geological Sciences, Rutgers University. Butler, R.F., 1992. Paleomagnetism: magnetic domains to geologic terranes. Blackwell Scientific Publications, Boston. Butler, R.F., Krause, D.W., Gingerich, P.D., 1987. Magnetic polarity stratigraphy and biostratigraphy of middle-late Paleocene continental deposits in south-central Montana. J. Geol. 95, 647–657. Carpenter, S.J., Erickson, M.J., Hoganson, J.W., Ludvigson, G.A., 2003. Alpine glaciation in the Laramide Mountains during the Late Cretaceous and Early Paleocene. Abstr. Prog. - Geol. Soc. Amer. 35, 291. Cerling, T.E., Quade, J., 1993. Stable carbon and oxygen isotopes in soil carbonates. In: Swart, P.K., Lohmann, K.C., McKenzie, J., Savin, S. (Eds.), Climate change in continental isotopic records. InAmerican Geophysical Union, Washington D.C, pp. 217–231. Cherven, V.B., Jacob, A.F., 1985. Evolution of Paleogene depositional systems, Williston Basin, in response to global sea level changes. In: Flores, R.M., Kaplan, S.S. (Eds.), Cenozoic paleogeography of the west-central United States. In: Rocky Mountain Paleogeography Symposium, vol. 3. Rocky Mountain Section, Society of Economic Paleontologists and Mineralogists, pp. 127–170. Cochran, J.K., Landman, N.H., Turekian, K.K., Michard, A., Schrag, D.P., 2003. Paleoceanography of the Late Cretaceous (Maastrichtian) Western Interior Seaway of North America: evidence from Sr and O isotopes. Palaeogeogr. Palaeoclimatol. Palaeoecol. 191, 45–64. Constenius, K.N., 1996. Late Paleogene extensionalcollapse of the Cordilleran foreland fold and thrust belt. Geol. Soc. Am. Bull. 108, 20–39. Coplen, T.B., Kendall, C., 2000. Stable hydrogen and oxygen isotope ratios for selected sites of the U.S. Geological Survey's NASQAN and benchmark surface-water networks. U.S. Geological Survery Open-File Report 00-160. Coughlin, T.J., O'Sullivan, P.B., Kohn, B.P., Holcombe, R.J., 1998. Apatite fission-track thermochronology of the Sierras Pampeanas central western Argentina: implications for the mechanism of plateau uplift in the Andes. Geology 26, 999–1002. Crowley, P.D., Reiners, P.W., Reuter, J.M., Kaye, G.D., 2002. Laramide exhumation of the Bighorn Mountains, Wyoming: an apatite (U–Th)/He thermochronology study. Geology 30, 27–30. Dansgaard, W., 1953. The abundance of 18O in atmospheric water and water vapour. Tellus 5, 461–469. DeCelles, P.D., 2004. Late Jurassic to Eocene evolution of the Cordilleran thrust belt and foreland basin system, western U.S.A. Am. J. Sci. 304, 105–168. DeCelles, P.D., Coogan, J.C., 2006. Regional structure and kinematic history of the Sevier fold-and-thrust belt, central Utah. Geol. Soc. Amer. Bull. 118, 841–864. DeCelles, P.D., Gray, M.B., Ridgway, K.D., Cole, R.B., Pivnik, D.A., Pequera, N., Srivastava, P., 1991. Controls on synorogenic alluvial-fan architecture, Beartooth Conglomerate (Paleocene), Wyoming and Montana. Sedimentology 38, 567–590. DeCelles, P.D., Quade, J., Kapp, P., Fan, M., Dettman, D.L., Ding, L., 2007. High and dry in central Tibet during the Late Oligocene. Earth Planet. Sci. Lett. 253, 389–401. Dettman, D.L., Lohmann, K.C., 1995. Microsampling carbonates for stable isotope and minor element analysis: physical separation of samples on a 20 micrometer scale. J. Sed. Res. 65, 566–569. Dettman, D.L., Lohmann, K.C., 2000. Oxygen isotope evidence for high-altitude snow in the Laramide Rocky Mountains of North America during the Late Cretaceous and Paleogene. Geology 28, 243–246. Dettman, D.L., Reische, A.K., Lohmann, K.C., 1999. Controls on the stable isotope composition of seasonal growth bands in aragonitic fresh-water bivalves (Unionidae). Geochim. Cosmochim. Acta 63, 1049–1057. Dickinson, W.R., Snyder, W.S., 1978. Plate tectonics of the Laramide orogeny. In: Matthews, V. (Ed.), Laramide folding associated with basement block faulting in the western United States: Geological Society of America Memoir, vol. 151, pp. 355–366. Dickinson, W.R., Klute, M.A., Hayes, M.J., Janecke, S.U., Lundin, E.R., McKittrick, M.A., Olivares, M.D., 1988. Paleogeographic and paleotectonic setting of Laramide sedimentary basins in the central Rocky Mountain region. Geol. Soc. Amer. Bull. 100, 1023–1039. Dutton, A., Wilkinson, B.H., Welker, J.M., Bowen, G.J., Lohmann, K.C., 2005. Spatial distribution and seasonal variation in 18O/16O of modern precipitation and river water across the conterminous USA. Hydrol. Process. 19, 4121–4146. Eberth, D.A., Hamblin, A.P., 1993. Tectonic, stratigraphic, and sedimentological significance of a regional discontinuity in the upper Judith River Group (Belly River Wedge) of southern Alberta, Saskatchewan, and northern Montana. Can. J. Earth Sci. 30, 174–200. Erslev, E.A., 1993. Thrusts, back-thrusts, and detachment of Rocky Mountain foreland arches. In: Schmidt, C.J., Chase, R., Erslev, E.A. (Eds.), Laramide basement deformation in the Rocky Mountain foreland of the western United States: Geological Society of America Special Paper, vol. 280, pp. 339–358. Erslev, E.A., 2005. 2D Laramide geometries and kinematics of the Rocky Mountains. In: Karlstrom, K.E., Keller, G.R. (Eds.), The Rocky Mountains region: an evolving lithosphere, tectonics, geochemistry, and geophysic. InAmerican Geophysical Union, Washington D.C, pp. 7–20. Fastovsky, D.E., 1987. Paleoenvironments of vertebrate-bearing strata during the Cretaceous-Paleocene transition, eastern Montana and western North Dakota. Palaios 2, 282–295. Flores, R.M., Ethridge, F.G., 1985. Evolution of intermontane fluvial systems of Tertiary Powder River Basin, Montana and Wyoming. In: Flores, R.M., Kaplan, S.S. (Eds.), Cenozoic paleogeography of the west-central United States. In: Rocky Mountain Paleogeography Symposium, vol. 3. Rocky Mountain Section, Society of Economic Paleontologists and Mineralogists, pp. 107–126. Fricke, H.C., 2003. Investigation of Early Eocene water-vapor transport and paleoelevation using oxygen isotope data from geographically widespread mammal remains. Geol. Soc. Amer. Bull. 115, 1088–1096. Garzione, C.N., Molnar, P., Libarkin, J.C., MacFadden, B.J., 2006. Rapid late Miocene rise of the Bolivian Altiplano: evidence for removal of mantle lithosphere. Earth Planet. Sci. Lett. 241, 543–556. Golovneva, L.B., 2000. The Maastrichtian (late Cretaceous) climate in the Northern Hemisphere. Geol. Soc. Lond., Spec. Publ. 181, 43–54. Gregory, K.M., Chase, C.G., 1992. Tectonic significance of plaeobotanically estimated climate and altitude of the late Eocene erosion surface, Colorado. Geology 20, 581–585. Gries, R., 1983. North-south compression of Rocky Mountain foreland structures. In: Lowell, J.D. (Ed.), Rocky Mountain Foreland Basins and Uplifts. InRocky Mountain Association of Geologists, Denver, Colorado, pp. 9–32. Grossman, E.L., Ku, T.L., 1986. Oxygen and carbon isotope fractionation in biogenic aragonite: temperature effects. Chem. Geol. 59, 59–74. Horton, T.W., Chamberlain, C.P., Fantle, M., Blum, J.D., 1999. Chemical weathering and lithologic controls of water chemistry in a high-elevation river system: Clark's Fork of the Yellowstone River, Wyoming and Montana. Water Resour. Res. 35, 1643–1655. Hoy, R.G., Ridgeway, K.D., 1997. Structural and sedimentological development of footwall growth synclines along an intraforeland uplift, east-central Bighorn Mountains, Wyoming. Geol. Soc. Am. Bull. 109, 915–935. Jordan, T.E., Allmendinger, R.W., 1986. The Sierras Pampeanas of Argentina: a modern analogue of Rocky Mountain foreland deformation. Am. J. Sci. 286, 737–764. Kent-Corson, M.L., Sherman, L.S., Mulch, A., Chamberlain, C.P., 2006. Cenozoic topographic and climatic response to changing tectonic boundary conditions in Western North America. Earth Planet. Sci. Lett. 252, 453–466. Kim, S., O'Neil, J.R., 1997. Equilibrium and nonequilibrium oxygen isotope effects in synthetic carbonates. Geochim. Cosmochim. Acta 61, 3461–3475. Koch, P.L., Zazhos, J.C., Dettman, D.L., 1995. Stable isotope stratigraphy and paleoclimatology of the Paleogene Bighorn Basin (Wyoming, USA). Palaeogeogr. Palaeoclimatol. Palaeoecol. 155, 61–89. Kohn, M.J., Dettman, D.L., 2007. Paleoaltimetry from stable isotope compositions of fossils. In: Kohn, M.J. (Ed.), Paleoaltimetry: geochemical and thermodynamic approaches: Reviews in Mineralogy & geochemistry, vol. 66, pp. 119–148. Lerbekmo, J.C., Coulter, K.C., 1985. Late Cretaceous to Early Tertiary magnetostratigraphy of a continental sequence: Red Deer Valley, Alberta, Canada. Can. J. Earth Sci. 22, 567–583. Lund, S.P., Hartman, J.H., Banerjee, S.K., 2002. Magnetostratigraphy of interfingering upper Cretaceous–Paleocene marine and continental strata of the Williston Basin, North Dakota and Montana. In: Hartman, J.H., Johnson, J.R., Nichols, D.J. (Eds.), the Hell Creek Formation and the Cretaceous–Tertiary boundary in the M. Fan, D.L. Dettman / Earth and Planetary Science Letters 286 (2009) 110–121 northern Great Plains: a integrated continental record of the end of the Cretaceous: Boulder, Colorado: Geological Society of America Special Paper, vol. 361, pp. 57–74. Mack, G.H., Cole, D.R., 2005. Geochemical model of δ18O of pedogenic calcite versus latitude and its application to Cretaceous palaeoclimate. Sed. Geol. 174, 115–122. McQuarrie, N., Chase, C.G., 2001. Raising the Colorado Plateau. Geology 28, 91–94. Morrill, C., Koch, P.L., 2002. Elevation and alteration? Evaluation of isotopic constraints on paleoaltitudes surrounding the Eocene Green River Basin. Geology 30, 151–154. Norris, R.D., Jones, L.S., Corfield, R.M., Cartlidge, J.E., 1996. Skiing in the Eocene Uinta Mountains? Isotopic evidence in the Green River Formation for snow melt and large mountains. Geology 24, 403–406. Obsborn, G., Stoclmal, G., Haspel, R., 2006. Emergence of the Canadian Rockies and adjacent plains: a comparison of physiography between end-of-Laramide time and the present day. Geomorphology 75, 450–477. Poage, M.A., Chamberlain, C.P., 2001. Empirical relationships between elevation and the stable isotope composition of precipitation and surface waters: considerations for studies of paleoelevation change. Am. J. Sci. 301, 1–15. Quade, J., Garzione, C.N., Eiler, J., 2007. Paleoelevation reconstruction using pedogenic. Carbonates Rev. Mineral. Geochem. 66, 53–87. Roberts, L.N.R., Kirschbaum, M.A., 1995. Paleogeography of the Late Cretaceous of the Western Interior of middle North America–coal distribution and sediment accumulation. U.S. Geol. Surv. Prof. Paper 1561, 115. Robinson, L.N., Honey, J.G., 1987. Geologic setting of a New Paleocene mammal locality in the northern Powder River Basin, Montana. Palaios 2, 87–90. Rowley, D.B., 2007. Stable isotope-based paleoaltimetry: theory and validation. Rev. Mineral. Geochem. 66, 23–52. Rowley, D.B., Pierrehumbert, R.T., Currie, B.S., 2001. A new approach to stable isotopebased paleoaltimetry: implications for paleoaltimetry and paleohypsometry of the High Himalaya since the Late Miocene. Earth Planet. Sci. Lett. 188, 253–268. Saleeby, J., 2003. Segmentation of the Laramide Slab-evidence from the southern Sierra Nevada region. Geol. Soc. Am. Bull. 115, 665–668. Secord, R., Gingerich, P.D., Smith, M.E., Clyde, W.C., Wilf, P., Singer, B.S., 2006. Geochrnology and mammalian biostratigraphy of middle and upper Paleocene continental strata, Bighorn Basin, Wyoming. Am. J. Sci. 306, 211–245. 121 Sewall, J.O., Sloan, L.C., 2006. Come a little bit closer: a high-resolution climate study of the Early Paleogene Laramide foreland. Geology 34, 81–84. Sklenar, S.E., Anderson, D.W., 1985. Origin and Early evolution of an Eocene lake system within the Washakie Basin of southwestern Wyoming. In: Reynolds, M.W., Dolly, E. D. (Eds.), Mesozoic Paleogeography of the West-central United States. In: Rocky Mountain Paleogeography Symposium, vol. 2. SEPM, pp. 231–246. Smith, M.E., Singer, B., Carroll, A., 2003. 40Ar/39Ar geochronology of the Eocene Green River Formation, Wyoming. Geol. Soc. Am. Bull. 115, 549–565. Strecker, U., 1996. Studies of Cenozoic Continental Tectonics. Part I. Apatite fission-track thermochronology of the Black Hills Uplift, South Dakota. PhD Dissertation, University of Wyoming, Laramie, Wyoming. Swisher III, C.C., Dingus, L., Butler, R.F., 1993. 40Ar/39Ar dating and magnetostratigraphic correlation of the terrestrial Cretaceous–Paleogene boundary and Puercan Mammal Age, Hell Creek–Tullock formations, eastern Montana. Can. J. Earth Sci. 30, 1981–1996. Tikoff, B., Maxson, J., 2001. Lithospheric buckling of the Laramide foreland during Late Cretaceous and Paleogene, western United States. Rocky Mount. Geol. 36, 13–35. Van der Pluijm, B.A., Vrolijk, P.J., David, R.P., Chris, M.H., Solum, J., 2006. Fault dating in the Canadian Rocky Mountains: evidence for late Cretaceous and Early Eocene orogenic pulses. Geology 34, 837–840. Wilf, P., 2000. Late Paleocene–Early Eocene climate changes in southwestern Wyoming: paleobotanical analysis. Geol. Soc. Amer. Bull. 112, 292–307. Wing, S.L., Bown, T.M., Obradovich, J.D., 1991. Early Eocene biotic and climatic change in interior western North America. Geology 19, 1189–1192. Wolberg, D.L., 1979. Late Paleocene (Tiffanian) mammalian fauna of two localities in eastern Montana. Northw. Geol. 8, 82–93. Wolfe, J.A., Forest, C.E., Molnar, P., 1998. Paleobotanical evidence of Eocene and Oligocene paleoaltitudes in midlatitude western North America. Geol. Soc. Amer. Bull. 110, 664–678. Zachos, J.C., Stott, L.D., Lohmann, K.C., 1994. Evolution of marine temperatures during the Palaeogene. Paleoceanography 9, 353–387. Zachos, J.C., Pagani, M., Sloan, L., Thomas, E., Billup, K., 2001. Trends, rhythms, and aberrations in global climate 65 Ma to present. Science 292, 686–693.
© Copyright 2025 Paperzz